Catalytic Enantioselective Synthesis of Tetrahydrofurans: A Dynamic

Feb 16, 2009 - Andrew T. Parsons and Jeffrey S. Johnson*. Department of Chemistry, University of North Carolina at Chapel Hill, Chapel Hill, North Car...
0 downloads 0 Views 150KB Size
Published on Web 02/16/2009

Catalytic Enantioselective Synthesis of Tetrahydrofurans: A Dynamic Kinetic Asymmetric [3 + 2] Cycloaddition of Racemic Cyclopropanes and Aldehydes Andrew T. Parsons and Jeffrey S. Johnson* Department of Chemistry, UniVersity of North Carolina at Chapel Hill, Chapel Hill, North Carolina 27599-3290 Received December 18, 2008; E-mail: [email protected]

Herein we report an efficient preparation of enantioenriched tetrahydrofuran (THF) derivatives through a dynamic kinetic asymmetric transformation (DyKAT)1 of racemic malonate-derived donor-acceptor (D-A) cyclopropanes via asymmetric [3 + 2] cycloaddition with aldehydes (eq 1).

D-A cyclopropanes have gained attention in the past decade because of their synthetic utility and ease of preparation.2 Recent literature has demonstrated that they provide access to carbocycles3 and heterocycles4,5 via rearrangement or [3 + 2], [3 + 3], and [3 + 4] cycloadditions with appropriate dipolarophiles, typically under Lewis acid catalysis. To date, generation of enantioenriched cycloadducts has typically required the use of nonracemic cyclopropane starting materials in reactions proceeding via stereospecific pathways.4,5k,n,6,7 Aldehydes undergo stereospecific Lewis acid-catalyzed cycloaddition with D-A cyclopropanes via a configurationally stable intimate ion pair to afford enantioenriched THF derivatives in high yields (eq 2).4,8 The background racemization of (S)-1a is

comparatively slow under these conditions.4,5k The identification of a kinetically competent racemization pathway was presumed to be a minimal precondition for the development of the more desirable transformation: a DyKAT of racemic D-A cyclopropanes through a chiral Lewis acid-catalyzed cycloaddition with aldehydes (eq 1). To this end, we sought to discover a Lewis acid that (1) selectively catalyzes the cycloaddition of the aldehyde with one enantiomer of cyclopropane and (2) promotes interconversion of the cyclopropane enantiomers. p-Methoxyphenylcyclopropane 1b is known to undergo rapid ring opening and was selected as the test substrate.4b The occurrence of a number of 2,5-diaryl-THF natural products bearing electronrich aromatic moieties further recommended 1b as an appropriate point of departure for this study.9 (tBu-pybox)MgI2 was identified as an effective catalyst for the cycloaddition of 1b and benzaldehyde, providing promising yield and enantiocontrol in the production of tetrahydrofuran 2b as a single diastereomer (Table 1, entry 1).10 Examination of other pybox ligands revealed that the tertbutyl group was necessary for sufficient selectivity and yield, and thus, we explored perturbations of tBu-pybox ligands at the 3122

9

J. AM. CHEM. SOC. 2009, 131, 3122–3123

Table 1. Examination of 4-Substituted tBu-pybox Ligandsa

entry

ligand

conversion (%)c

yield (%)c

erd

1 2 3 4 5 6 7 8 9f 10g 11

L1 L2 L3 L4 L5 L6 L7 L8 L8 L8 L9

93 26 100 100 95 100 100 100 100 100 100

62 5 57 73 59 67 55 74e 16 40 75

95.5:4.5 nd 96.5:3.5 95.5:4.5 94:6 96:4 93.5:6.5 96:4 82.5:17.5 96:4 95:5

a Conditions: 1b (1.0 equiv), benzaldehyde (2.0 equiv), MgI2 (0.10 equiv), L (0.12 equiv), [1b]0 ) 0.05 M in CCl4, room temperature, 48 h. b Determined by 1H NMR spectroscopy of the unpurified product. c Determined by 1H NMR spectroscopy using a mesitylene internal standard. d Determined by chiral SFC analysis. e Average isolated yield of two trials. f With CH2Cl2 as the solvent. g With C7H8 as the solvent.

4-position of the pyridine. Altering the substituent X of the 4-Xt Bu-pybox ligands (L2-L9) had a negligible effect on enantioselectivity but induced a significant variation in yield. Ligands L8 (X ) Cl) and L9 (X ) Br) provided the highest yields and levels of enantiocontrol (Table 1, entries 8 and 11). L8 · MgI2 catalyzed the enantioselective cycloaddition of various aldehydes with rac-1b (Table 2). This reaction worked well for electron-rich aryl, cinnamyl, and linear and branched aliphatic aldehydes. Electron-poor aldehydes typically gave lower yields, presumably because of their poor nucleophilicity.4 This methodology was also extended to activated cyclopropanes bearing 2-thienyl (1c) and (E)-CHdCHPh (1d) donor groups. The remaining mass balance was attributed to cyclopropane decomposition, affording a complex mixture of byproducts. The cycloaddition of rac-1a with p-anisaldehyde catalyzed by L8 · MgI2 yielded 2p with an er of 95.5:4.5; the product was determined to be the (R,R) enantiomer by comparison to previously reported SFC retention times and optical rotation.4 The recovered cyclopropane was found to be enriched in the (R) enantiomer (er ) 86:14; Table 3, line 1), and therefore, rac-1a is a substrate for a simple kinetic resolution rather than a DyKAT.7 This result suggests that the aldehyde selectively reacts with the (S) enantiomer 10.1021/ja809873u CCC: $40.75  2009 American Chemical Society

COMMUNICATIONS a-c

Table 2. Substrate Scope

(S,S) configuration. Conversely, use of (S)-1a affords optically pure (R,R)-2p in excellent yield with near complete conversion of the cyclopropane (Table 3). In summary, a simple protocol for the preparation of enantioenriched THF derivatives through a dynamic kinetic asymmetric [3 + 2] cycloaddition of racemic malonate-derived D-A cyclopropanes and aldehydes has been developed. A variety of cyclopropanes bearing electron-rich donor groups undergo cycloaddition with aryl, cinnamyl, and aliphatic aldehydes to afford products in good yields and enantioselectivity. Current efforts seek to elucidate the mechanism of this new transformation, understand the unique reactivity of the (pybox)MgI2 catalysts, and extend this methodology to other asymmetric reactions with D-A cyclopropanes. Acknowledgment. This work was supported by the NSF (CHE0749691). The authors thank Professor Michel Gagne´ for helpful discussions and Mr. Joe Bailey for assistance in ligand synthesis. Supporting Information Available: Experimental details and characterization data for new compounds. This material is available free of charge via the Internet at http://pubs.acs.org. References

a Conditions: 1 (1.0 equiv), aldehyde (2.0-4.0 equiv), MgI2 (0.10 equiv), L8 (0.11 equiv), [1]0 ) 0.05 M in CCl4, room temperature, 7-56 h. See the Supporting Information for additional experimental details. b Yield refers to the average isolated yield of at least two trials. c er was determined by chiral SFC analysis. d Determined by 1H NMR spectroscopy of the unpurified product. e dr > 25:1. f er was determined for the major (cis) diastereomer.

Table 3. Stereochemical Analysisa

substrate

conversion (%)b

yield (%)c

(R,R)-2p/ (S,S)-2pd,e

(R)-1a/ (S)-1af

rac-1a (R)-1a (S)-1a

54 22 96

38 3 92

95.5:4.5 21:79 99.5:0.5

86:14 98:2 25:75

a Conditions: see Table 2, footnote a. b The difference in yield and conversion is attributed to cyclopropane decomposition. c The yield of 2p was determined by 1H NMR spectroscopy using a mesitylene internal standard and is the average of two trials. d dr > 99:1. e er was determined by chiral SFC analysis. f er of recovered 1a as determined by chiral GC analysis.

of the cyclopropane through a stereospecific nucleophilic attack.4,5l This interpretation is corroborated by the sluggish reaction of enantiopure (R)-1a, which gives product 2p with er ) 79:21 in the

(1) Faber, K. Chem.sEur. J. 2001, 7, 5004–5010. (2) For D-A cyclopropane reviews, see: (a) Reissig, H.-U. Top. Curr. Chem. 1988, 144, 73–135. (b) Wong, H. N. C.; Hon, M. Y.; Tse, C. W.; Yip, Y. C.; Tanko, J.; Hudlicky, T. Chem. ReV. 1989, 89, 165–198. (c) Reissig, H.-U.; Zimmer, R. Chem. ReV. 2003, 103, 1151–1196. (d) Yu, M.; Pagenkopf, B. L. Tetrahedron 2005, 61, 321–347. (3) (a) Shimizu, I.; Ohashi, Y.; Tsuji, J. Tetrahedron Lett. 1985, 26, 3825– 3828. (b) Hiroi, K.; Yamada, A. Tetrahedron: Asymmetry 2000, 11, 1835– 1841. (c) Korotkov, V. S.; Larionov, O. V.; de Meijere, A. Synthesis 2006, 3542–3546. (d) Ivanova, O. A.; Budynina, E. M.; Grishin, Y. K.; Trushkov, I. G.; Vereletskii, P. V. Angew. Chem., Int. Ed. 2008, 47, 1107–1110. (4) (a) Pohlhaus, P. D.; Johnson, J. S. J. Am. Chem. Soc. 2005, 127, 16014– 16015. (b) Pohlhaus, P. D.; Sanders, S. D.; Parsons, A. T.; Li, W.; Johnson, J. S. J. Am. Chem. Soc. 2008, 130, 8642–8650. (5) (a) Sugita, Y.; Kawai, K.; Yokoe, I. Heterocycles 2000, 53, 657–664. (b) Sugita, Y.; Kawai, K.; Yokoe, I. Heterocycles 2001, 55, 135–144. (c) England, D. B.; Kuss, T. D. O.; Keddy, R. G.; Kerr, M. A. J. Org. Chem. 2001, 66, 4704–4709. (d) Young, I. S.; Kerr, M. A. Angew. Chem., Int. Ed. 2003, 42, 3023–3026. (e) Ganton, M. D.; Kerr, M. A. J. Org. Chem. 2004, 69, 8554–8557. (f) Pohlhaus, P. D.; Johnson, J. S. J. Org. Chem. 2005, 70, 1057–1059. (g) Cardona, F.; Goti, A. Angew. Chem., Int. Ed. 2005, 44, 7832–7835. (h) Kang, Y.-B.; Tang, Y.; Sun, X.-L. Org. Biomol. Chem. 2006, 4, 299–301. (i) Bowman, R. K.; Johnson, J. S. Org. Lett. 2006, 8, 573–576. (j) Korotkov, V. S.; Larionov, O. V.; Hofmeister, A.; Magull, J.; de Meijere, A. J. Org. Chem. 2007, 72, 7504–7510. (k) Sapeta, K.; Kerr, M. A. J. Org. Chem. 2007, 72, 8597–8599. (l) Karadeolian, A.; Kerr, M. A. J. Org. Chem. 2007, 72, 10251–10253. (m) Bajtos, B.; Yu, M.; Zhao, H. D.; Pagenkopf, B. L. J. Am. Chem. Soc. 2007, 129, 9631– 9634. (n) Perreault, C.; Goudreau, S. R.; Zimmer, L. E.; Charette, A. B. Org. Lett. 2008, 10, 689–692. (o) Parsons, A. T.; Campbell, M. J.; Johnson, J. S. Org. Lett. 2008, 10, 2541–2544. (6) (a) Sibi and coworkers reported an enantioselective synthesis of cis- and trans-tetrahydro-1,2-oxazines via cycloaddition of racemic D-A cyclopropanes and nitrones, perhaps resulting from enantiodifferentiation after the first irreversible step. See: Sibi, M. P.; Ma, Z.; Jasperse, C. P. J. Am. Chem. Soc. 2005, 127, 5764–5765. For a definition and discussion, see: (b) Zhang, W.; Lee, N. H.; Jacobsen, E. N. J. Am. Chem. Soc. 1994, 116, 425–426. (7) Tang and co-workers reported a simple kinetic resolution of racemic DA cyclopropanes to access enantioenriched tetrahydro-1,2-oxazines and cyclopropanes via cycloaddition with nitrones. See: Kang, Y.-B.; Sun, X.L.; Tang, Y. Angew. Chem., Int. Ed. 2007, 46, 3918–3921. (8) For a review of stereoselective THF synthesis, see: Wolfe, J. P.; Hay, M. B. Tetrahedron 2007, 63, 261–290. (9) Ward, R. S. Nat. Prod. Rep. 1999, 16, 75–96. (10) For use of MgI2 as a catalyst for cycloadditions involving cyclopropanes, see: (a) Alper, P. B.; Meyers, C.; Lerchner, A.; Siegel, D. R.; Carreira, E. M. Angew. Chem., Int. Ed. 1999, 38, 3186–3189. (b) Lautens, M.; Han, W. J. Am. Chem. Soc. 2002, 124, 6312–6316. (c) Bertozzi, F.; Gustafsson, M.; Olsson, R. Org. Lett. 2002, 4, 3147–3150. (d) Ref 5e. For (pybox) MgX2 complexes as stoichiometric enantioselective promoters, see: (e) Yamamoto, H.; Hayashi, S.; Kubo, M.; Harada, M.; Hasegawa, M.; Noguchi, M.; Sumimoto, M.; Hori, K. Eur. J. Org. Chem. 2007, 2859–2864.

JA809873U

J. AM. CHEM. SOC.

9

VOL. 131, NO. 9, 2009 3123